Rajdeep Mukherjeea,
Susanta Banerjee*a,
Hartmut Komberb and
Brigitte Voitb
aMaterials Science Centre, Indian Institute of Technology, Kharagpur 721302, India. E-mail: susanta@matsc.iitkgp.ernet.in; Fax: +91-3222-255303; Tel: +91-3222-283972
bLeibniz-Institut für Polymerforschung Dresden e.V., Hohe Strasse 6, 01067, Dresden, Germany
First published on 27th August 2014
A series of new copolymers HPPQSH-XX PS were synthesized from preformed sulfonated ionomers HPPQSH-XX in order to get high proton conductivity along with other excellent properties. They were prepared by direct copolymerization of bisphenol HPP with two bishalides, QBF and SDCDPS. Again the copolymers were post-sulfonated using 1,3-propanesultone in the presence of NaH and were analysed by spectroscopic techniques. The random copolymers HPPQSH-XX with different statistical distribution of –SO3H moiety showed very small ionic clusters (5–10 nm) whereas the grafted copolymers showed larger ionic domains (60–100 nm) in their TEM images. All ionomer membranes showed good mechanical properties, high oxidative and dimensional stability with low water uptake and swelling ratios. Ion exchange capacities (weight and volume based) were also calculated to explain better the correlation with water uptake and proton conductivity (14–125 mS cm−1 at 80 °C and 15–142 mS cm−1 at 90 °C under fully hydrated condition) of the membranes.
Several classes of advanced aromatic hydrocarbon-based sulfonated polymers were synthesized to develop suitable PEM materials such as poly(arylene ether)s,4,5 poly(arylene ether ketone)s,6–8 poly(aryl ether ether ketone ketone)s,9–11 poly(arylene ether sulfone)s,12–19 poly(arylene sulfide sulfone)s,20,21 polyimides,22,23 and polytriazoles.24–26 Depending upon their structural features, each class of polymers exhibited a unique set of properties in fuel technology. However, as a class of high-performance engineering thermoplastic materials, poly(arylene ether)s have some excellent properties like high glass transition temperature, high thermal and chemical stability along with good mechanical properties. In order to get some superior PEM properties, functionalization of a polymer is essential for modifying the chemical structure of the polymer.
Generally, two functionalization techniques are considered to functionalize poly(arylene ether)s. The first one is direct polymerization or copolymerization by polycondensation between the functionalized monomers and the sulfonated monomers; the second one is the post-sulfonation technique, where the preformed polymer with no sulfonic acid group (–SO3H) is sulfonated by using different modifier or by applying sulfonated functionalized reagents. However, in the direct polymerization method, the sulfonic acid groups are usually located in the polymer backbone and those polymers generally show acceptable conductivities only at high ion exchange capacities which finally results in extensive water uptake above a critical degree of sulfonation or a critical temperature (percolation threshold) along with a dramatic loss of mechanical properties. Thus the post-sulfonation technique was considered to overcome such limitations by modifying the polymers. However, to enhance the mechanical integrity along with distinct separation of hydrophilic moiety from hydrophobic segment in a polymer matrix, the side chain sulfonation method has attracted much attention. Besides, in post-sulfonation, choosing the right monomer and control of grafting percentage are crucial to prepare the desired product.
Perhaps for the first time, Jannasch et al. reported a post-sulfonation method where polysulfone (PSU) was sulfonated by lithiation followed by the reaction of the lithiated sites with 2-sulfobenzoic acid cyclic anhydride (SBACA).27 Later on the same group also reported a grafting method where PSU was initially reacted with butyllithium to form an organolithium compound which was further reacted with 4-fluorobenzoyl chloride followed by functionalized sulfonated monomers to introduce the sulfonic acid group in the polymer side chain.28 Watanabe and co-workers reported the synthesis of fluorenylbiphenyl group-containing poly(arylene ether)s followed by post-sulfonation using chlorosulfonic acid to improve the hydrolytic stability by incorporating the sulfonic acid group in the side chain and keeping the polymer main chain in a hydrophobic environment.29 For the first time Lin et al. developed novel main-chain-type and side-chain-type sulfonated poly(ether ether ketone) membranes with a single polymer backbone by reacting the sulfonic acid groups of pristine SPEEKs with 2-aminoethanesulfonic acid to improve the nanophase-separated morphology of the material.30 Guiver et al. synthesized aromatic poly(arylene ether sulfone) copolymers by using a new fluorinated bis-phenol monomer followed by the incorporation of sulfonic acid group by choosing a suitable sulfonated phenolic monomer. These grafted sulfonated copolymers showed excellent thermal and chemical stability. In addition, Guiver and co-workers also synthesized rigid aromatic comb-shaped poly(arylene ether sulfone) copolymers wherein the sulfonic acid sites are on linear or branched pendant chains. They reported that such polymers with the sulfonated groups attached to pendent side groups are very stable under heat, hydrolysis and oxidation.31,32 Recently, Guiver et al. reported a novel class of fully aromatic comb-shaped polymers with highly sulfonated aromatic graft chains that can self-assemble into nanoscale organized structures and display high proton conductivity over a wide range of humidity along with better stability.33 So, in order to get better PEM materials with desired set of properties still more work is required.
In the present work, a series of high molecular weight fluorinated sulfonated poly(arylene ether sulfone) copolymers (schematically shown in Scheme 1) were prepared by applying a combination of the direct polymerization or copolymerization method followed by the post-sulfonation technique. At first, HPPQS-XX copolymers with varying degree of sulfonation are prepared using direct polycondensation of 4,4′-bis(4′-fluoro-3′-trifluoromethylbenzyl)biphenyl (QBF) and 3,3′-disodiumsulfonyl-4,4′-dichlorodiphenylsulfone (SDCDPS) with 3,3′-bis(4-hydroxyphenyl)-1-isobenzopyrrolidone (HPP) to incorporate the sulfonic acid group into the polymer backbone. Finally, HPPQS-XX PS copolymers are synthesized from the preformed HPPQS-XX copolymers by applying the post-sulfonation technique using 1,3-propanesultone in the presence of sodium hydride as a base. In order to get high mechanical strength, good phase-separated morphology and high proton conductivity with low swelling, a sulfonated aliphatic side chain was introduced with proper tuning of hydrophobic and hydrophilic moieties in the polymer matrix. Thus the present work demonstrates the feasibility and success of this synthetic methodology along with detailed investigation of the morphology and selected PEM properties such as proton conductivity, water uptake, mechanical strength, swelling ratio, and thermal and chemical stability.
![]() | ||
Scheme 1 Representation of rigid polyaromatic sulfonated backbone with linear pendant side chains containing sulfonic acid groups. |
A variety of PEMs have been prepared either by the direct copolymerization method with one or two pendant sulfoalkyl groups as a side chain37–39 or by chemically grafting the pendants onto polymers.40–45 In this regard, the direct copolymerization technique is the more efficient method to control the degree of sulfonation in polymers rather than the chemical grafting method. Except for a few cases, preparation of most sulfonated monomers with high content of sulfonic groups requires extremely rigorous conditions, stringent moisture-free environment and painstaking efforts. Thus it is very difficult to synthesize those sulfonated monomers in purified form. The chemical grafting method is the only way to avoid the synthesis of sulfonated monomers and offers an easier route to introduce the sulfonic moiety as a pendant side chain. Gieselman and Reynolds introduced a sulfopropyl group to the amide nitrogen of aramid poly(p-phenyleneterephthalamide) (PPTA) through a nucleophilic ring-opening reaction with 1,3-propanesultone using NaH as a base.46 By applying this method, Gao et al. synthesized a series of cardo poly(arylene ether sulfone)s with different lengths containing pendant sulfoalkyl groups as a side chain which are attached to the nitrogen centre of the lactam ring.41 It has been found that, for primary and secondary amides, N-alkylation is more favoured than O-alkylation (except when the counterion is Ag+) if the reaction was carried out at room temperature using a strong base, but the reverse phenomenon was observed at elevated temperature probably due to the higher reactivity of the O-centre. So in order to get the desired N-sulfopropylated product we treated preformed sulfonated polymers (HPPQS-XX) with 1,3-propanesultone using a strong base, NaH, at room temperature for 12 h (Scheme 2). Finally, the viscous reaction mixture was precipitated in isopropanol, and the product washed with distilled water, filtered and dried under vacuum at 100 °C for at least 24 h. Repeatedly (2–3 times) the post-sulfonation technique was carried out either by adding a greater amount of NaH or 1,3-propanesultone or by adding both to get 100% post-sulfonation, but unfortunately complete post-sulfonation (maximum 83% side chain sulfonation) did not occur in either case, as was evident from NMR spectra. GPC results (Table 1) showed the formation of high molecular weight copolymers (before and after post-sulfonation) with relatively narrow dispersities (Đ). However, the values as presented in Table 1 are not a true representation of the molar mass of the polymers. These polymers (polyelectrolytes) interact partly with the GPC column and the separation was not according to hydrodynamic volume (size) as was observed from GPC. Thus, the molecular weight and polydispersity values of the copolymers did not show any direct relationship with copolymer composition.
Polymer | Mna | Đb | DS | |
---|---|---|---|---|
Theo.c | NMRd | |||
a Mn, number-average molecular weight.b Đ, dispersity.c Degree of sulfonation is theoretically calculated from the monomer feed ratio. For PS samples the value represents complete alkylation.d Calculated from 1H NMR signal intensities. | ||||
HPPQ | 48![]() |
2.11 | 0 | 0 |
HPPQSH-10 | 21![]() |
1.74 | 0.1 | 0.06 |
HPPQSH-20 | 19![]() |
1.60 | 0.2 | 0.17 |
HPPQSH-30 | 22![]() |
1.74 | 0.3 | 0.24 |
HPPQSH-PS | 58![]() |
2.14 | 1.0 | 0.73 |
HPPQSH-10 PS | 24![]() |
1.59 | 1.1 | 0.88 |
HPPQSH-20 PS | 19![]() |
1.58 | 1.2 | 1.00 |
HPPQSH-30 PS | 23![]() |
1.79 | 1.3 | 1.06 |
The chemical structures of all copolymers (HPPQS-XX and HPPQS-XX PS) were confirmed by ATR-FTIR, 1H NMR, 13C NMR and 19F NMR spectroscopy. The ATR-FTIR spectra of homopolymer HPPQ and its post-sulfonated form HPPQS-PS are shown in Fig. S1a.† The characteristic symmetric stretching band at 1051 cm−1 clearly indicated the presence of aromatic ether linkages in the polymers. In the spectrum of the post-sulfonated polymer (HPPQS PS) an additional absorption band appeared at 1039 cm−1 corresponding to the stretching vibration of aliphatic sulfonic acid group. Besides, the strong absorption band at 1700 cm−1 was shifted to 1676 cm−1 (corresponding to the stretching vibration of secondary amide carbonyl and tertiary amide carbonyl respectively) indicating the formation of tertiary nitrogen centre after post-sulfonation. The FTIR spectra of copolymers are shown in Fig. S1b.† The characteristic symmetric stretching band at 1051 cm−1 and asymmetric stretching bands at 1333 and 1483 cm−1 were attributed to the aromatic ether linkages in the polymers. In addition, both homopolymer and copolymers exhibited aromatic CC stretching band at 1584 cm−1 and C–F stretching band at 1240–1127 cm−1. All the copolymers (before and after grafting) showed symmetric and asymmetric stretching bands of aromatic sodium sulfonate at 1027 cm−1 and 1095 cm−1. These two absorption bands increased gradually with an increase of SDCDPS content. Additionally, in the spectra of post-sulfonated copolymers two characteristic stretching bands appeared at 1039 cm−1 and 1676 cm−1 corresponding to the aliphatic sulfonic acid group and tertiary amide carbonyl, respectively. The peaks associated with aromatic sulfonic acid group were shifted towards higher frequency with simultaneous increase of DS which indicated the possibility of hydrogen bonding with bound water. This phenomenon becomes more significant in post-sulfonated copolymers due to the incorporation of additional –SO3H groups as grafted side chains.
Fig. 1a depicts the 1H NMR spectrum of HPPQ. The analysis of this spectrum and of the corresponding 13C NMR spectrum (Fig. S2†) is straightforward and confirms the expected structure. The signal assignment for the post-sulfonated polymer HPPQSH PS is more challenging because the alkylation of the NH group is not complete and results in a terpolymer structure. The 1H NMR spectrum (Fig. 1b) is characterized by new alkyl signals (a–c) resulting from the formed N-(3-sulfopropyl) group but also by residual NH signal indicating incomplete conversion with 1,3-propanesultone. The new signals in the aromatic proton and carbon regions (Fig. S3†) were assigned by combination of 1D and 2D NMR techniques. Complete 1H and 13C signal assignments are given in the ESI.† The degree of alkylation can be calculated from the intensities of NH and alkyl signals. However, signal overlap with signals from (H2O + –SO3H), from DMSO-d6 and/or from residual solvent from synthesis hampers accurate integration. Thus, the intensity ratios NH0/HAr(HPPQ) for the parent HPPQ and NHPS/HAr(HPPQSH PS) for the PS sample were determined (Fig. 1). Because the number of aromatic protons remains unchanged after post-sulfonation reaction, the degree of alkylation can be calculated according to eqn (1):
Degree of alkylation = 1 − [(NHPS × HAr(HPPQ))/(NH0 × HAr(HPPQSH PS))]. | (1) |
Replacing quadriphenyl (Q) units partially by sulfonated diphenylsulfone (SDPS) units results in the copolymer series HPPQSH-XX. The alternating structure of HPPQ changes to a random terpolymer with a characteristic microstructure. Typically, the 1H and 13C NMR signals of the 4-oxyphenyl moiety of the phthalimidine unit show a dyad splitting due to the different neighboring units (Q or SDPS).35 In fact, such a splitting can be observed both in the 1H and in the 13C NMR spectra of HPPQSH-XX samples (Fig. 2a–c and 3). Additionally, the NH signal of the HPP unit shows a splitting due to the three different HPP centered triads (Fig. 2a–c). Relating the content of the three triads to the polymer composition, a good agreement with a random distribution of Q and SDPS units can be stated. The copolymer composition itself was calculated from integral values of regions S (H26 of SDPS), Q (H15 and H17 of Q), and P (H3 of HPP). It was found for all samples that the SDPS content is lower than expected from the monomer feed (Table 1). The appearance of weak signals at 6.74 ppm resulting from ortho protons of phenolic groups of HPP and complete conversion of fluorine groups as proved by 19F NMR analysis point to a slight excess of HPP monomer in the final copolymer. Whereas the 1H and 13C NMR spectra of HPPQSH-XX samples could be completely assigned (see also ESI†), the NMR spectra after partial alkylation are too complex (Fig. 2d–f). Thus, only the degree of alkylation was calculated according to eqn (1) using the signal integrals of NH protons and all aromatic protons of the parent HPPQSH-XX sample and the corresponding PS sample. The degree of alkylation was 82 (±2)% for all samples.
All the acid-form copolymer membranes including their post-sulfonated analogues showed two-step degradation profiles in TGA (Fig. 4). The 10% degradation temperatures are listed in Table 2. The first weight loss was found at around 240–300 °C, which corresponds to the degradation of labile sulfonic acid moiety. However, these initial weight loss values gradually decrease with increasing DS. The second degradation temperature was observed in the range of 450–530 °C and corresponds to the decomposition of the polymer backbone.
Polymer | Tgb (°C) | Td10%c (°C) | TSd (MPa) | Ye (GPa) | EBf (%) | Oxidative stabilityg (h) | |
---|---|---|---|---|---|---|---|
τ1 | τ2 | ||||||
a Data in parentheses refer to the salt (–Na) form.b Glass transition temperature determined by DSC, heating rate 10 °C min−1 under nitrogen.c 10% degradation temperature measured by TGA, heating rate 10 °C min−1 under nitrogen atmosphere.d Tensile strength, at a strain rate of 5% min−1, 65 ± 2% RH and 30 °C.e Young's modulus.f Elongation at break.g τ1 and τ2 refer to the initial breaking time and the time to complete dissolution in Fenton's reagent (2 ppm FeSO4 in 3% H2O2) at 80 °C respectively. | |||||||
HPPQ | 302 | 524 | 96 | 2.37 | 56 | — | |
HPPQSH-10 | 310 | 466 | 61 (78) | 2.52 (2.15) | 6 (10) | >24 | — |
HPPQSH-20 | 322 | 373 | 58 (65) | 2.03 (1.79) | 4 (11) | >24 | — |
HPPQSH-30 | — | 317 | 38 (40) | 1.42 (1.45) | 4 (6) | >24 | — |
HPPQSH-PS | — | 310 | 70 (72) | 1.97 (1.89) | 20 (36) | >24 | — |
HPPQSH-10 PS | — | 285 | 69 (52) | 2.20 (1.67) | 9 (7) | 22 | >24 |
HPPQSH-20 PS | — | 263 | 65 (51) | 2.55 (1.64) | 4 (4) | 10.2 | >24 |
HPPQSH-30 PS | — | 245 | 37 (29) | 1.35 (1.01) | 4 (12) | 5.5 | 15.2 |
Nafion® 117 | — | — | 21.9 | 0.16 | 288 | — | — |
Mechanical stability is one of the essential requirements for electrolyte membranes under extreme conditions for membrane electrode assemblies (MEAs) to be used in fuel cells. The synthesized copolymers exhibited very good mechanical properties in both salt and acid forms under dry state at room temperature, and the results are summarized in Table 2 together with data for Nafion® 117. The change of mechanical behaviour was also revealed by considering the stress–strain plots shown in Fig. 5. It is clear that the tensile strength of the salt form of copolymers HPPQS-XX (XX = 10, 20, 30) gradually decreased with an increase of DS due to simultaneous increase of aromatic –SO3H groups in the polymer backbone and was found to be in the range of 40–78 MPa with Young's modulus of 1.45–2.15 GPa. Similar observation was also made for the post-sulfonated copolymers HPPQS-XX PS with tensile strength varying within 29–72 MPa, Young's modulus of 1.01–1.89 GPa, and elongation at break of 4–36%. The HPPQSH-XX and HPPQSH-XX PS membranes (acid form) had tensile strength in the range of 37–70 MPa, Young's modulus of 1.35–2.55 GPa, and elongation at break of 4–20%. The acid form of the post-sulfonated copolymer (HPPQSH-XX PS) membranes showed better mechanical properties than their corresponding salt form as can be seen from Table 2. This was somewhat contrary to literature findings.17,36 Possibly, the sulfonated flexible aliphatic side chains in post-sulfonated copolymers (HPPQSH-XX PS) participated in intermolecular hydrogen bonding with backbone –SO3H groups as well as the unreacted –NH groups as depicted schematically in Fig. 6. Such type of protonic cross linking was not feasible in the salt-form membranes due to the unavailability of H+ ions, and in the case of backbone-sulfonated rigid HPPQSH-XX copolymers due to steric congestion. Compared to the soft fluorocarbon-based Nafion® 117 these copolymers showed higher tensile strength and Young's modulus but lower elongation, attributed to the presence of rigid aromatic backbone structure (rigid quadriphenyl and bulky phthalimidine moieties).
The oxidative stability of HPPQSH-XX and HPPQSH-XX PS copolymers was evaluated by observing the breaking time (τ1) of the membranes in Fenton's reagent (3 wt% H2O2, 2 ppm FeSO4) at 80 °C. All the polymer films exhibited high oxidative stability, as shown in Table 2. The trend of oxidative stability was quite similar to that of our previous work;36 it decreased with increasing degree of sulfonation of the membranes. Naturally, in this series, a maximum oxidative stability (τ1 > 24 h) was found for the HPPQSH-10 copolymer whereas the minimum oxidative stability (τ1 = 5.5 h and τ2 = 15.2 h) was found for the HPPQSH-30 PS copolymer. Compared to our previous set of copolymers, these post-sulfonated membranes showed higher oxidative stability even at higher IEC value.36 In fact, oxidative attack mainly occurs at the hydrophilic domains of the polymer. But in post-sulfonated membranes the oxidative stability is highly influenced by the following three factors – the first is the wholly aromatic nature of the copolymers, next is the presence of sulfonic acid group in already electron-deficient aromatic sulfonic acid moieties, and finally the aliphatic side chains separate the polymer main chain from hydrophilic sulfonic acid groups to keep the main chain in a hydrophobic surrounding. These three factors together minimize the degradation of polymer chains during oxidative attack. In addition, these membranes showed higher oxidative stability due to the presence of hydrophobic trifluoromethyl groups which can protect the polymer backbone from attack of hydroperoxy and hydroxyl radicals.23
The hydrolytic stability of all membranes (before and after post-sulfonation) was investigated by immersing the membranes in water at 100 °C for 24 h. No obvious changes were observed on the basis of their appearance, weight, and IECW value, which indicates an excellent hydrolytic behaviour of all synthesized copolymers.
Polymer | dMa (g cm−3) | IECW (meq. g−1) | IECVd (meq. g−1) | Water uptakee (wt%) | Water uptakef (vol%) | |||||
---|---|---|---|---|---|---|---|---|---|---|
dry | Wet | |||||||||
Theo.b | Titr. | NMRc | 30 °C | 80 °C | 90 °C | 90 °C | 90 °C | |||
a Density of membrane calculated from the weight and dimension of dry sample.b IECW,Theo. = (1000/MWrepeat unit) × DSTheo. × 2, where DSTheo. is calculated theoretically from monomer feed ratio.c IECW,NMR = (1000/MWrepeat unit) × DSNMR × 2, where DSNMR is calculated from NMR peak ratio.d IECV(dry) = (IECW,Theo.) × dM and IECV(wet) = IECV(dry)/(1 + 0.01WU).e WU (wt%) = [(Wwet − Wdry)/Wdry] × 100.f WU (vol%) = [{(Wwet − Wdry)/dw}/(Wdry/dm)] × 100. | ||||||||||
HPPQSH-10 | 2.12 | 0.27 | 0.20 | 0.16 | 0.57 | 0.44 | 0.44 | 0.43 | 14 | 30 |
HPPQSH-20 | 2.02 | 0.54 | 0.51 | 0.46 | 1.09 | 0.85 | 0.83 | 0.81 | 17 | 34 |
HPPQSH-30 | 1.81 | 0.81 | 0.72 | 0.65 | 1.47 | 1.13 | 1.09 | 1.08 | 21 | 37 |
HPPQSH-PS | 1.91 | 1.14 | 0.98 | 0.83 | 2.18 | 1.61 | 1.54 | 1.52 | 23 | 44 |
HPPQSH-10 PS | 1.88 | 1.38 | 1.13 | 1.08 | 2.59 | 1.88 | 1.73 | 1.68 | 29 | 54 |
HPPQSH-20 PS | 1.61 | 1.62 | 1.49 | 1.35 | 2.61 | 1.85 | 1.71 | 1.63 | 37 | 59 |
HPPQSH-30 PS | 1.64 | 1.86 | 1.67 | 1.51 | 3.06 | 2.05 | 1.90 | 1.81 | 42 | 69 |
Nafion® 117 | 1.96 | 0.91 | 0.90 | — | 1.77 | 1.29 | 1.10 | — | — | — |
Sulfonated polymers with ideal water uptake (WU) value along with acceptable mechanical properties is one of the critical demands for their application as PEMs. In order to facilitate a more exact, precise discussion about the water uptake among different membranes, volumetric ion exchange capacity (IECV, meq. cm−3) is more relevant (Table 3) which is defined as the molar concentration of sulfonic acid groups per unit volume containing absorbed water. In addition, as the electrochemical properties like proton conductivity are directly related with the length scale (independent of mass), more reasonable explanations are obtained when IECV (dependent on length) is considered rather than IECW. The water uptake [WU (wt%) and WU (vol%)] of copolymer membranes (before and after grafting) was measured at various temperatures of 30 °C, 80 °C and 90 °C. Fig. 7 represents the variation of IECW against WU (wt%) of the ionomer membranes at the three different temperatures. As shown in Table 3, WU (wt%) increases with increasing sulfonation degree at a given temperature; thus HPPQSH-30 PS showed the highest WU (wt%) of 37% at 80 °C and 42% at 90 °C among all the copolymer membranes. Compared to Nafion® 117 membrane, this new set of copolymers showed much lower value of WU (wt%) at all temperatures except HPPQSH-30 PS which may be attributed to the higher IECW value of HPPQSH-30 PS compared to that of Nafion® 117 membrane. This new set of copolymers HPPQSH-XX and post-sulfonated form HPPQSH-XX PS showed also much lower WU (wt%) values than our previous set of main-chain-type sulfonated copolymers36 (maximum 44% for PAQSH-60 at 80 °C) even at higher degree of sulfonation which indicates the hydrophobic nature of the aliphatic group as a side chain in the grafted copolymers. Additionally, the lower WU (wt%) values of the copolymers were also influenced by the presence of a hydrophobic –CF3 group in the quadriphenyl moiety and rigid aromatic polymer backbone. The volume-based water uptake, WU (vol%), was calculated according to previous reports.19,43 The density values of the polymers are listed in Table 3 and were used to calculate IECV of the dry membranes. The variation of dry- and wet-based IECV against WU (vol%) of the ionomer membranes is shown in Fig. 8. At lower temperature (30 °C) the change of dry volume-based IECV and wet volume-based IECV was quite similar when plotted against WU (vol%) of the polymers with different DS values. But at higher temperature (80 °C) IECV (wet) showed wide variation from IECV (dry) probably due to the percolation effect. Such effect was more prominent after reaching a certain IECV (wet) value on raising the temperature to 90 °C.
![]() | ||
Fig. 7 Water uptake (wt%) as a function of IECW and temperature for HPPQSH-XX and HPPQSH-XX PS copolymers. |
![]() | ||
Fig. 8 Water uptake (vol%) dependence of IECV (dry) and IECV (wet) values of HPPQSH-XX and HPPQSH-XX PS membranes. |
Swelling ratio of the membranes increased with increasing DS and temperature, as expected. Fig. 9 represents the in-plane (length) and the through-plane (thickness) dimensional change of different copolymers at fixed temperature together with data for Nafion® 117. The results are listed in Table 4. Compared to the many main chain type copolymers, the side chain ionomer membranes generally showed much lower dimensional swelling due to the presence of flexible hydrophobic alkyl groups as side chains which are more effective in repelling polar water molecules.6,7,9,42 Similar results were found for HPPQSH-XX PS copolymers where the grafted membranes showed much lower swelling ratio than main chain type sulfonated copolymers. In addition, the in-plane dimensional swelling ratio was much lower than the through-plane dimensional swelling ratio.43
Polymer | Swelling ratio (%) | σa (mS cm−1) | Eab (kJ mol−1) | |
---|---|---|---|---|
In plane | Through plane | |||
90 °C | 90 °C | 90 °C | ||
a Proton conductivity was measured under fully hydrated condition (in water).b Activation energy determined in the temperature range 30–90 °C and heating rate 1–2 K min−1. | ||||
HPPQSH-10 | 3 | 21 | 15 | 10.6 |
HPPQSH-20 | 5 | 24 | 24 | 14.8 |
HPPQSH-30 | 8 | 28 | 27 | 12.5 |
HPPQSH-PS | 10 | 33 | 32 | 14.9 |
HPPQSH-10 PS | 13 | 36 | 45 | 12.1 |
HPPQSH-20 PS | 18 | 45 | 90 | 12.4 |
HPPQSH-30 PS | 22 | 48 | 142 | 13.9 |
Nafion® 117 | — | — | 150 | 13.6 |
This observation is well explained on considering the TEM analysis of the copolymer membranes. The TEM images (Fig. 11) show an outstanding microphase-separated morphology in all cases where the dark spherical regions correspond to the hydrophilic ionic domains (lead ion exchange –SO3H group) and the brighter regions represent the hydrophobic domains. The distinct ionic domains were more interconnected after post-sulfonation of the copolymers with gradual increase of ion exchange capacities. Only the preformed sulfonated backbone-containing copolymers (HPPQSH-XX, XX = 10, 20, 30) consist of a large amount of small ionic clusters (5–10 nm) that were dispersed throughout the hydrophobic polymer matrix. The particle sizes were greatly increased after post-sulfonation of the copolymer membranes indicating the possibility of higher agglomeration of hydrophilic moieties due to insertion of additional pendant –SO3H groups. Thus the grafted copolymers exhibited excellent phase-separated morphology with greater extent of bigger ionic domains (60–100 nm) along with fewer medium ionic clusters (15–20 nm) which can provide better water channel along with good proton transport pathway. Compared to our previous set of copolymers PAQSH-XX (containing only linear backbone –SO3H groups), this new set of post-sulfonated copolymers HPPQSH-XX PS (containing –SO3H groups both in the linear backbone and in pendant side chains) exhibited much better phase-separated morphology probably due to the presence of sulfoalkyl groups as a side chain that facilitated an additional interaction with the main chain sulfonic acid groups.
![]() | ||
Fig. 12 Proton conductivity of all copolymer membranes at different temperatures under fully hydrated conditions. |
![]() | ||
Fig. 13 Arrhenius-type temperature-dependent proton conductivity (σ) behaviour of all copolymer membranes along with Nafion® 117. |
The activation energy of copolymer membranes was also calculated and was found to be in the range of 14.9–10.6 kJ mol−1. In this regard, the activation energy (13.9 kJ mol−1) of HPPQSH-30 PS membranes was quite close to the value of Nafion® 117 (13.6 kJ mol−1) which allowed to assume an analogous type of proton conduction mechanism involving hydronium ions. Also, there may be a possibility of more effective interaction between two polymer chains due to the formation of protonic crosslinking involving intermolecular hydrogen bonding. It can be observed from Table 4 that the activation energy values of the copolymers did not show any change, but rather increased as the DS value increased for some of the copolymers. This is somewhat difficult to explain as there are many factors that affect the activation energy. However, as the temperature increases, the protonic crosslinking involving intermolecular hydrogen bonding breaks and that could be a reason for no change of activation energy in the threshold of percolation (between HPPQSH-10 PS and -20 PS). It could be simply stated that the copolymers have activation energies in the range of that of Nafion® 117. Furthermore, a correlation plot of proton conductivity with IECW and IECV (wet%) of all ionomer membranes is shown in Fig. 14. At low IECW values (up to 1.38 meq. g−1), the hydrophilic clusters are far apart from each other that leads to a poor interconnection and lower proton conductivity. Whereas at higher IECW (1.62 meq. g−1 and greater), the isolated hydrophilic clusters come closer to build a well-connected channel by formation of a large ionic cluster (validated by the microstructural evidence from TEM analysis) which results in high proton conductivity.
![]() | ||
Fig. 14 Correlation plot of proton conductivity with IECW and IECV (wet%) of all ionomer membranes at 80 °C. |
Footnote |
† Electronic supplementary information (ESI) available: FT-IR spectra, 1H, 13C & 19F NMR data and spectra, DSC plot of the HPPQSH-XX membranes, Nyquist plot of ionomer membranes. See DOI: 10.1039/c4ra07291j |
This journal is © The Royal Society of Chemistry 2014 |